Fact-checked by Grok 2 weeks ago

Delocalized electron

In chemistry, delocalized electrons are electrons that are not confined to a single atom or localized but instead spread across multiple atoms within a , , polyatomic structure, or solid material, often through mechanisms like or orbital overlap. This delocalization typically involves π electrons in conjugated systems or free-moving electrons in metallic lattices, leading to enhanced molecular and unique physical properties such as electrical conductivity. Unlike localized electrons, which are associated with specific bonds or atoms, delocalized electrons distribute charge and energy more evenly, lowering the overall energy of the system. Delocalized electrons play a central role in covalent and . In organic and inorganic molecules with conjugated π systems, such as (C₆H₆), electrons from alternating double bonds are delocalized across the ring, resulting in equal bond lengths of 139 pm—intermediate between typical single (154 pm) and double (134 pm) C–C bonds—and conferring aromatic stability through energy. Similarly, in (O₃), the π electrons delocalize between structures, yielding identical O–O bonds of 127.2 pm, which blend single (148 pm) and double (120.7 pm) characteristics. In metals, the electron sea model describes delocalized valence electrons surrounding a of positive ions; for instance, in sodium, the 3s electrons move freely, enabling the nondirectional bonding that underlies metallic properties. The presence of delocalized electrons imparts significant functional advantages. In conjugated polymers like (PANI) or poly(3,4-ethylenedioxythiophene) (PEDOT), these electrons facilitate along the chain, supporting applications in conductive materials and biosensors. In metals such as magnesium, greater delocalization of electrons strengthens compared to sodium, correlating with higher melting points (650°C versus 97.8°C) and improved thermal conductivity, as the mobile electrons efficiently transfer heat by colliding with lattice ions. Overall, delocalization stabilizes reactive intermediates like carbocations and explains phenomena from color in dyes to plasmonic effects in metal nanoparticles.

Fundamentals

Definition

Delocalized electrons are electrons in a , , or that are not associated with a single atom or a specific but instead occupy an extended orbital spread over several to many atoms. This delocalization typically arises in systems such as conjugated π-electron networks or conduction bands in solids, where the electrons are distributed across a or multiple atomic centers rather than being confined to discrete positions. A primary characteristic of delocalized electrons is their enhanced mobility, enabling them to traverse the structure more freely than localized electrons. Additionally, delocalization contributes to the of chemical systems by reducing the overall through the spreading of , making the structure more favorable than equivalent localized arrangements. In quantum mechanical terms, these electrons are depicted by wavefunctions that span extended regions, reflecting their probabilistic distribution over multiple atoms. Unlike localized electrons, which are tightly associated with particular orbitals or bonds—such as the σ electrons in alkanes—delocalized electrons allow for shared that supports stabilizing effects like . This distinction highlights how delocalization promotes greater flexibility in electron arrangement compared to the more rigid localization in isolated bonds. At the quantum mechanical foundation, delocalized electrons are described by wavefunctions that satisfy the across the entire relevant region of the system, such as a or , thereby capturing their non-localized nature through solutions to the time-independent . This approach underscores that the behavior of these electrons emerges from the eigenvalues and eigenfunctions of the system's total energy operator, emphasizing their extended spatial extent.

Historical Development

The concept of delocalized electrons emerged in the early 20th century as quantum mechanics provided a framework for understanding electron behavior beyond localized atomic orbitals. In 1926, Erwin Schrödinger introduced his wave equation, which described electrons as wave functions rather than particles, laying the groundwork for non-localized descriptions of electron distribution in atoms and molecules. Concurrently, Werner Heisenberg's matrix mechanics formulation in 1925 emphasized quantum uncertainty, influencing later ideas of electron resonance and delocalization. Building on these foundations, Linus Pauling advanced valence bond theory in the 1930s, introducing the resonance concept to account for electron delocalization in molecules like benzene, where electrons are shared across multiple bond structures rather than fixed positions. Pauling's seminal 1931 paper formalized this approach, demonstrating how resonance hybrid structures lower energy through delocalized π-electrons. Key milestones in the 1920s and 1930s further solidified the delocalized electron paradigm. initiated in 1927–1928, proposing that electrons occupy orbitals spanning entire molecules, as seen in his analysis of diatomic spectra. Robert S. Mulliken expanded this in the early 1930s through the method, emphasizing delocalized "spectroscopic" molecular orbitals that explained bonding and spectra in polyatomic systems. In , Felix Bloch's 1928 theorem described electron wave functions in periodic crystal lattices as modulated plane waves, enabling the band theory that portrayed conduction s as delocalized across the material. These developments shifted from isolated atomic models to extended electron distributions. In , the 1930s saw targeted applications to π-systems. Erich Hückel applied in 1931 to derive his for aromatic , predicting that planar cyclic systems with 4n + 2 delocalized π-electrons exhibit enhanced stability due to closed-shell configurations. Complementing this, Erich Clar developed his notation in the mid-20th century for representing delocalized electrons in polycyclic aromatic hydrocarbons, using circles to denote aromatic sextets and emphasizing migratory π-electron character over fixed bonds. Post-World War II advancements integrated these ideas into broader contexts, particularly metals. Paul Drude's 1900 treated conduction electrons as a , explaining basic electrical properties empirically. refined this quantum mechanically in 1927, incorporating Fermi-Dirac statistics to describe delocalized electrons filling bands up to the , resolving classical inconsistencies like specific . By the late 1940s and 1950s, these concepts merged with in , fostering band theory's application to semiconductors and metals, marking the transition from molecular to .

Theoretical Frameworks

Valence Bond Theory and Resonance

Valence bond theory describes chemical bonding through the overlap of atomic orbitals from individual atoms, forming localized bonds that share electron pairs. In this framework, atoms often undergo , where atomic orbitals of similar energy combine to form hybrid orbitals better suited for bonding; for instance, carbon in molecules typically uses sp² hybridization to create three σ bonds in a planar arrangement, leaving a p orbital for π bonding. Delocalization arises when a single cannot adequately represent the electron distribution, leading to the use of multiple contributors that average to a , effectively spreading electrons over several atoms rather than localizing them between pairs. The concept in represents delocalized electrons, particularly in π systems, by drawing canonical structures that depict alternative placements of electrons while maintaining the same atomic positions. For example, is described by two primary Kekulé structures, each showing three alternating double bonds, but the actual exists as a of these forms, resulting in equivalent bonds throughout the ring. This superposition accounts for the delocalization of the six π electrons across the entire ring, stabilizing the system beyond what a single structure predicts. Resonance energy quantifies the stabilization from this delocalization and is calculated as the difference between the energy of the most stable reference structure and that of the resonance hybrid: \Delta E_{\text{res}} = E_{\text{reference}} - E_{\text{hybrid}} For benzene, this value is approximately 36 kcal/mol, indicating significant lowering of the ground-state energy compared to a hypothetical localized cyclohexatriene. In representations of such systems, delocalized π electrons are often shown with dashed lines or circles encompassing the affected bonds to emphasize their non-localized nature; experimental evidence includes the equidistant C-C bond lengths in benzene, measured at 1.39 Å, intermediate between typical single (1.54 Å) and double (1.34 Å) bonds. Despite its utility for molecular π systems, remains qualitative and struggles with extensive delocalization, as it relies on discrete resonance structures rather than continuous electron distributions, making it less effective for describing in where electrons are highly delocalized across lattices.

provides a quantum mechanical framework for describing the electronic structure of , where electrons occupy (MOs) formed as linear combinations of atomic orbitals (LCAOs). In this approach, the wave function of a is expressed as ψ_MO = ∑ c_i φ_i, where φ_i are the atomic orbitals and c_i are coefficients determined by solving the variationally. These MOs are inherently delocalized, extending over the entire rather than being confined to specific bonds, and they occur in and antibonding pairs that determine molecular stability. This delocalization arises because the electrons respond to the potential of all nuclei simultaneously, leading to wave functions that spread across multiple atoms. Delocalization is particularly evident in the occupation of π or σ molecular orbitals, where electrons in antibonding π* or bonding σ orbitals contribute to the overall electronic spanning the molecular framework. For extended systems, such as those with higher , -adapted linear combinations (SALCs) of orbitals are employed to construct that transform according to the irreducible representations of the molecular , ensuring proper matching and facilitating delocalization over symmetric frameworks. This method highlights how delocalized s in π systems, for instance, lower the energy compared to localized descriptions by allowing greater freedom in electron . In small molecules like (CH₄), delocalized illustrate partial delocalization even in saturated systems. The four bonding are formed from the carbon 2s and 2p orbitals combining with the four 1s orbitals, resulting in tetrahedral σ where is distributed symmetrically across the C-H bonds, though not strictly localized to individual pairs. These show high between carbon and atoms but extend over the whole molecule, demonstrating how LCAO approximations capture the shared nature of electrons in polyatomic structures. Computationally, methods such as Hartree-Fock theory solve for these delocalized orbitals using the Roothaan equations, which reformulate the Hartree-Fock problem in an LCAO basis to iteratively optimize the coefficients c_i and orbital energies. The resulting canonical MOs are delocalized by nature, but for interpretive purposes, they can be transformed into localized orbitals using methods like the Boys procedure, which maximizes the sum of squared distances between orbital charge centroids, or the Pipek-Mezey approach, which localizes based on atomic population projections. These transformations aid in bridging quantum calculations with classical bonding concepts without altering the underlying delocalized electron description.

Molecular Examples

Conjugated Systems

Conjugated systems in molecules feature alternating single and double bonds, enabling the delocalization of π electrons across a chain of sp²-hybridized carbon atoms. This structure arises from the overlap of adjacent p orbitals perpendicular to the molecular plane, forming extended π molecular orbitals that span multiple atoms rather than being localized to individual bonds. A classic example is 1,3-butadiene (CH₂=CH–CH=CH₂), where four p orbitals combine to produce four π molecular orbitals: two bonding (ψ₁ and ψ₂, occupied by four π electrons) and two antibonding (ψ₃ and ψ₄). The delocalization in these systems lowers the energy gap between the highest occupied (HOMO) and the lowest unoccupied (LUMO), with the gap decreasing as the conjugation length increases. In 1,3-butadiene, the HOMO (ψ₂) and LUMO (ψ₃) separation corresponds to a π → π* transition absorbing light at approximately 217 nm. Extended polyenes, such as those with additional double bonds, exhibit even smaller gaps, shifting absorption to longer wavelengths (e.g., 258 nm for 1,3,5-hexatriene). This effect enhances molecular stability, as evidenced by lower heats of for conjugated dienes (223–229 kJ/mol) compared to isolated alkenes (251 kJ/mol). Additionally, delocalization leads to bond length equalization: in polyenes, C=C bonds elongate to 1.34–1.37 and C–C bonds shorten to 1.42–1.44 , intermediate between typical single (1.48 ) and double (1.34 ) bond lengths, with the effect saturating in chains longer than 6–7 double bonds. These properties confer enhanced reactivity to conjugated systems, particularly in reactions like the Diels-Alder, where dienes adopt an s-cis conformation to react with dienophiles, forming six-membered rings with high . The delocalized electrons stabilize the , making conjugated dienes more reactive than isolated alkenes. absorption via π → π* transitions is a hallmark, with longer polyenes like (e.g., β-carotene, with 11 conjugated double bonds) displaying visible color due to absorption in the 400–500 nm range, appearing orange-red as they reflect longer wavelengths. Hückel molecular orbital theory provides a simple model for these systems, treating the π electrons as moving in a of p orbitals with nearest-neighbor interactions parameterized by α (Coulomb integral) and β (resonance integral). For a linear polyene with n p-orbital sites, the molecular orbital energies are given by E_j = \alpha + 2\beta \cos\left( \frac{\pi j}{n+1} \right) where j = 1, 2, ..., n, yielding bonding orbitals for lower j values and antibonding for higher ones. This approximation captures the delocalization and energy ordering observed in (n=4) and longer polyenes.

Aromatic Compounds

Aromatic compounds represent a class of cyclic, planar molecules featuring a continuous loop of conjugated p-orbitals where π electrons are delocalized, resulting in enhanced stability beyond that of typical conjugated systems. This delocalization occurs when the system satisfies , which states that a planar, monocyclic with 4n + 2 π electrons (where n is a non-negative ) exhibits aromatic character due to the formation of a closed-shell electronic configuration in the molecular orbitals. The delocalized π electrons form a uniform cloud above and below the ring plane, contributing to the characteristic properties of . The prototypical example is (C₆H₆), a six-membered ring with six π electrons delocalized evenly over the six carbon atoms. This delocalization is often depicted using a circle inscribed within the hexagon to symbolize the uniform , contrasting with localized bond representations. The aromatic stabilization, known as the aromatic sextet energy, amounts to approximately 150 kJ/mol, as determined from experiments comparing to hypothetical localized models. Structural evidence for this delocalization includes uniform C–C bond lengths of 1.39 Å—intermediate between typical single (1.54 Å) and double (1.34 Å) bonds—and bond angles of 120°, consistent with sp² hybridization and a planar D₆h . Beyond benzene, polycyclic systems like (C₁₀H₈) demonstrate extended delocalization with 10 π electrons across two fused rings, satisfying for n = 2 and yielding a stabilization energy of about 255 kJ/mol. Larger monocyclic annulenes, such as annulene (C₁₈H₁₈) with 18 π electrons (n = 4), also exhibit aromatic stability when planar, showing diatropic ring currents and low reactivity typical of delocalized systems. In contrast, systems with 4n π electrons, like cyclobutadiene (C₄H₄) with four π electrons (n = 1), are anti-aromatic; their forced planarity leads to destabilized delocalization, resulting in a distorted rectangular structure with alternation and high reactivity./15:_Benzene_and_Aromaticity/15.03:Aromaticity_and_the_Huckel_4n%2B_2_Rule) Spectroscopic techniques provide direct evidence of delocalized electrons in aromatic compounds through the phenomenon of ring currents. In (NMR) spectroscopy, the circulating delocalized π electrons generate a secondary that deshields protons on the ring periphery, shifting their signals upfield (e.g., protons at ~7.3 ppm) compared to non-aromatic analogs. This diatropic ring current effect, first quantified for aromatic systems, distinguishes aromatic delocalization from paratropic (anti-aromatic) behavior, where inner protons are shielded and outer ones deshielded.

Solid-State Examples

Metallic Bonding

In metallic bonding, metal atoms form a lattice of positively charged cations surrounded by a "sea" of delocalized valence electrons, primarily from the s and p orbitals, which are free to move throughout the structure. This electron sea model, first proposed by Paul Drude in the early 1900s, conceptualizes the valence electrons as detached from their parent atoms and shared collectively among all cations in the lattice. The delocalized nature of these electrons arises because the energy required to excite them into higher molecular orbitals is low, allowing them to occupy extended states across the entire metal crystal. The bonding mechanism relies on the electrostatic attraction between the fixed positive metal cations and the mobile cloud, which holds the together without forming discrete, directional bonds between individual atoms. This lack of fixed bonds permits the layers of cations to slide past one another under applied stress, as the electron sea acts as a , preventing and enabling the characteristic and malleability of metals. For instance, in ductile metals like , deformation allows realignment of the without breaking the cohesive electron-mediated interactions. Representative examples illustrate the role of delocalized electrons in . In sodium metal, each atom contributes its single 3s to the , resulting in one delocalized electron per atom and a relatively low of 97.8°C due to weaker overall attraction. Transition metals, such as iron or , involve contributions from both s/p and d electrons; for example, the 3d and 4s electrons in these elements delocalize, increasing the in the sea and leading to stronger and higher melting points, like iron's 1538°C. A quantum mechanical refinement of the classical treats the delocalized electrons as a gas, where electrons occupy quantized energy states in accordance with the . In this model, electrons fill energy levels up to the , the highest occupied state at , which determines the distribution and availability of electrons for bonding and transport; for typical metals, the ranges from 2 to 10 eV. This quantum description better accounts for the stability of the electron sea by incorporating wave-like behavior and Fermi-Dirac statistics, improving upon Drude's classical kinetic theory. The in this model provides evidence for , manifesting in properties such as high luster, where delocalized electrons reflect incident light efficiently, and superior thermal conductivity, as mobile electrons transfer rapidly through the . For example, metals like silver exhibit thermal conductivities around 400 W/m·K at , far exceeding those of non-metals, due to this electron-mediated .

Band Theory in Solids

Band theory describes the electronic of periodic solids, where delocalized s occupy energy bands formed by the overlap of orbitals across the . In a with a periodic potential, the yields solutions known as Bloch waves, which capture the wave-like propagation of s while respecting the periodicity. The foundational principle is the Bloch theorem, which states that the wavefunctions in a periodic potential can be expressed as \psi_{\mathbf{k}}(\mathbf{r}) = u_{\mathbf{k}}(\mathbf{r}) e^{i \mathbf{k} \cdot \mathbf{r}}, where u_{\mathbf{k}}(\mathbf{r}) is a periodic function matching the translation symmetry, and \mathbf{k} is the wavevector within the first Brillouin zone. This form implies that s are delocalized plane waves modulated by the , enabling their mobility throughout the solid. Energy bands arise when discrete orbitals from each site interact and split into a continuum of states, with the number of states in each band equaling the number of atoms times the degeneracy of the contributing orbitals. The band width depends on the orbital overlap strength, broader for tightly bound systems like metals. In the resulting band structure, the valence band—formed from filled atomic orbitals—is fully occupied at , while the conduction band, derived from empty or higher-energy orbitals, lies above it. These bands are separated by a E_g, the forbidden energy region where no states exist; E_g = 0 in metals (overlapping or partially filled bands), small (\sim 0.1-3 ) in semiconductors, and large (> 3 ) in insulators. Delocalization manifests in partially filled bands, where s near the can move freely under an , contributing to ; in empty conduction bands of semiconductors, or optical excitation promotes s across E_g, enabling limited mobility. For example, exhibits an indirect of approximately 1.1 at , allowing some delocalized carriers via thermal generation, whereas diamond's direct of about 5.5 results in highly localized s and insulating behavior. The Kronig-Penney model illustrates band formation mathematically in a one-dimensional lattice with periodic delta-function potentials, solving the time-independent Schrödinger equation to derive the dispersion relation \cos(ka) = \cos(\kappa a) + P \frac{\sin(\kappa a)}{\kappa a}, where k is the Bloch wavevector, a the lattice constant, \kappa = \sqrt{2mE}/\hbar, and P a strength parameter. Allowed energies occur where |\cos(ka)| \leq 1, forming bands separated by gaps at Brillouin zone edges (k = n\pi/a), demonstrating how periodicity splits free-electron levels into delocalized band states. This framework extends to non-metals like , where delocalized \pi electrons from carbon p_z orbitals form overlapping within basal planes, yielding a semimetallic with near-zero and high in-plane , as visualized in band diagrams showing linear dispersion near the Dirac points. Perpendicular to the planes, weaker interlayer localizes electrons, resulting in anisotropic .

Properties and Applications

Electrical Conductivity

Delocalized electrons are fundamental to in metals, where they form a "" of free charge carriers that respond to an applied by drifting through the . In the , proposed in , these are treated as classical particles undergoing random collisions with ions, leading to a mean relaxation time \tau between collisions. Under an E, the electrons acquire a v_d = -\frac{e \tau}{m} E, where e is the electron charge and m its , resulting in a j = -n e v_d./06:_Elements_of_Kinetics/6.02:_The_Ohm_law_and_the_Drude_formula) The resulting is given by \sigma = \frac{n e^2 \tau}{m}, where n is the ; this formula captures the high of metals, with typical values of \sigma around $10^7 S/m for at . In semiconductors, delocalized electrons contribute to through thermal across the or via doping, which introduces additional charge carriers. Intrinsic semiconductors have a small density of thermally excited electrons in the conduction and holes in the valence , enabling modest that increases exponentially with temperature as more carriers are promoted. Doping enhances this by substituting atoms: n-type doping (e.g., in ) adds donor impurities that release extra delocalized electrons to the conduction , while p-type doping (e.g., ) creates acceptors that leave mobile holes in the valence for conduction. These delocalized carriers can yield conductivities up to $10^3 S/m in heavily doped , far exceeding intrinsic values. Graphite exemplifies anisotropic arising from planar delocalization of \pi electrons within its layered structure, where each layer behaves like a two-dimensional metal. In-plane reaches approximately $10^5 S/m due to the free movement of these delocalized electrons parallel to the sheets, while perpendicular is orders of magnitude lower (\sim 10^2 S/m) because electron hopping between layers is limited by weak van der Waals interactions. In contrast, insulators like exhibit negligible because their electrons are tightly localized in covalent bonds, with no accessible delocalized states in the conduction band at ; 's wide of 5.5 prevents thermal excitation of carriers, resulting in resistivity exceeding $10^{14} \Omega \cdotm.

Optical Properties

Delocalized electrons in conjugated systems facilitate optical absorption through π→π* electronic transitions, where the excitation promotes an electron from the highest occupied () to the lowest unoccupied (LUMO). The extent of conjugation reduces the HOMO-LUMO energy gap, leading to absorption in the ; for instance, in polyenes like , longer conjugated chains result in lower-energy transitions and the characteristic orange-red coloration. This delocalization enhances the of these transitions, contributing to high absorption coefficients in organic materials. In metals, delocalized conduction electrons exhibit collective oscillations at the plasma frequency, defined as \omega_p = \sqrt{\frac{4\pi n e^2}{m}}, where n is the , e the electron charge, and m the . Electromagnetic waves with frequencies below \omega_p are strongly reflected due to the screening effect of these free electron plasmons, accounting for the metallic luster observed in the visible range. Above \omega_p, typically in the ultraviolet for most metals, penetration occurs, transitioning the material toward . Aromatic compounds with delocalized π electrons display symmetry-forbidden transitions that gain intensity through vibronic coupling, where vibrational modes distort the to allow otherwise prohibited excitations. This coupling enables emission from delocalized s, as seen in polycyclic aromatic hydrocarbons, where the radiative decay from the produces characteristic UV-visible . In solid-state semiconductors, delocalized electrons in band structures permit direct band-to-band transitions, particularly efficient in materials like (GaAs) with a direct bandgap of 1.42 eV, enabling light emission in LEDs via electron-hole recombination without assistance. Representative examples include synthetic dyes such as azo compounds, where extended π conjugation delocalizes electrons across the , inducing bathochromic shifts that tune from to visible wavelengths for applications in textiles and sensors. In organic photovoltaics, this delocalization in conjugated polymers enhances light harvesting by broadening spectra through reduced excitonic binding energies.

References

  1. [1]
    Delocalization of Electrons - Chemistry LibreTexts
    Jan 29, 2023 · Electron delocalization means electrons move around and are not static, shown by resonance forms, and is related to the movement of π electrons ...
  2. [2]
  3. [3]
    9.5: Metallic Bonding - Chemistry LibreTexts
    Jun 5, 2019 · A strong metallic bond will be the result of more delocalized electrons, which causes the effective nuclear charge on electrons on the cation to ...
  4. [4]
    delocalization of electrons (08789) - IUPAC
    A delocalized electron is not associated with a particular atom or particular covalent bond but occupies an extended orbital spread over several to many atoms ...Missing: definition | Show results with:definition
  5. [5]
    delocalization (D01583) - IUPAC Gold Book
    Delocalization, a quantum concept in organic chemistry, describes π in conjugated systems where bonding is not localized between two atoms.Missing: definition | Show results with:definition
  6. [6]
    Delocalized Bonding and Molecular Orbitals
    Molecular orbital theory is a delocalized bonding approach that explains the colors of compounds, their stability, and resonance.
  7. [7]
    [PDF] ELECTRON DELOCALIZATION AND RESONANCE
    Since electrons are charges, the presence of delocalized electrons brings extra stability to a system compared to a similar system where electrons are.
  8. [8]
    7.7 Molecular Orbital Theory – Chemistry Fundamentals
    Just like electrons around isolated atoms, electrons around atoms in molecules are limited to discrete (quantized) energies. The region of space in which a ...Learning Outcomes · Bonding In Diatomic... · The Diatomic Molecules Of...
  9. [9]
    Molecular Orbital theory (MO) is the most important ... - Chemistry 301
    The MO perspective on electrons in molecules is very different from that of a localized bonding picture such as valence bond (VB) theory. In VB we describe ...
  10. [10]
  11. [11]
    Wave Function - an overview | ScienceDirect Topics
    In this method it is recognized from the outset that in a molecule electrons are delocalized over the whole region of space spanned by the nuclear framework.
  12. [12]
  13. [13]
    [PDF] Robert S. Mulliken - Nobel Lecture
    Molecular orbitals also began to appear in a fairly clear light as suitable homes for electrons in molecules in the same way as atomic orbitals for electrons in ...
  14. [14]
    Introduction: DelocalizationPi and Sigma | Chemical Reviews
    It would be far better to follow Clar's recommendation to restrict the circle notation to the representation of six aromatic electrons. The fourteen reviews ...
  15. [15]
    [PDF] The development of the quantum-mechanical electron theory of metals
    The electron theory of metals underwent dramatic development between the first proposal by Paul Drude. (1900) and Hendrik Antoon Lorentz (1904–1905, 1909) at.
  16. [16]
    Valence Bond Theory and Resonance (M9Q4)
    This section explores valence bond theory and how it is applied to molecules with multiple bonds and those that need resonance structures to better describe ...
  17. [17]
    RESONANCE THEORY
    Resonance theory uses multiple valence structures, called resonance structures, to represent molecules with delocalized electrons, like benzene, where no ...
  18. [18]
  19. [19]
    [PDF] Valence Bond Theory
    Valence bond theory does not easily address delocalization. ... The sulfur atom is surrounded by two bonds and one lone pair of electrons in either resonance ...
  20. [20]
    Valence Bond and Molecular Orbital: Two Powerful Theories that ...
    Nov 18, 2021 · As the names suggest, valence bond theory considers a molecule as a collection of bonds, whereas molecular orbital theory views the molecule as ...Missing: formula | Show results with:formula
  21. [21]
    Are Molecular Orbitals Delocalized? | Journal of Chemical Education
    Mar 12, 2012 · The bonding electrons in methane can be properly described in terms of localized electrons, and photoelectron spectroscopy does not indicate otherwise.
  22. [22]
    Localized and delocalized molecular orbital description of methane
    The purpose of this article is to show that the relationship between localized and delocalized molecular orbitals can be easily demonstrated for the case of ...
  23. [23]
    Unravelling the Origin of Intermolecular Interactions Using ...
    In this paper we have presented an energy decomposition analysis (EDA) method based on absolutely localized molecular orbitals (ALMO) ...
  24. [24]
    None
    ### Summary of Conjugated Systems (Non-Aromatic) from Chapter 12
  25. [25]
  26. [26]
  27. [27]
    DFT studies on the structural and vibrational properties of polyenes
    Apr 6, 2016 · The change in terminal C = C bond length as a result of increasing the all-trans polyene chain length, expressed by the number of double bonds ( ...
  28. [28]
    Diels Alder Reaction: Dienes and Dienophiles - Chemistry Steps
    First, a reminder that the Diels-Alder reaction is a type of pericyclic reaction between a conjugated diene (two double bonds) and a dienophile (an alkene with ...
  29. [29]
    Mechanisms Underlying Carotenoid Absorption in Oxygenic ...
    We thus propose that these proteins are able to tune the absorption of their red-absorbing carotenoid via the rotation of conjugated end cycles toward the ...
  30. [30]
    [PDF] Lecture notes on Huckel Theory - MSU chemistry
    * Molecular orbitals have the form. 4. 4. I. 4 = Σ fici. L= 1. 5 four pe ... Linear Polyene: Consider a "linear" molecule with N carbons. C с. Са. C. N. C. 6 7.Missing: formula | Show results with:formula
  31. [31]
    [PDF] Linear Polyene Electronic - Structure and Potential Surfaces
    First, it is a con- strained refinement in which all of the C-H bond lengths and CC-H bond angles were assumed to be equal, and the molecular symmetry was.
  32. [32]
    Quantentheoretische Beiträge zum Benzolproblem
    Quantentheoretische Beiträge zum Benzolproblem. Download PDF. Erich Hückel. 3875 Accesses. 1887 Citations. 25 Altmetric. 5 Mentions. Explore all metrics ...
  33. [33]
    Benzene's bond lengths corrected | Research - Chemistry World
    Sep 22, 2022 · Benzene has six C-C bonds of equal distance (Pauling: Ref. 8 in [1]) of d(CC) = 1.39 Å (+/-0.02), which was shown (see Fig.1 in [1]) to be ...
  34. [34]
  35. [35]
  36. [36]
    Structures of Metals
    A reasonable model would be one in which atoms are held together by strong, but delocalized, bonds. Bonding. Such bonds could be formed between metal atoms that ...
  37. [37]
  38. [38]
    [PDF] Introduction to Solid State Physics
    This book is the eighth edition of an elementary text on solid state/ condensed matter physics for seniors and beginning graduate students of the physical ...
  39. [39]
    [PDF] Über die Quantenmechanik der Elektronen in Kristallgittern
    Von Felix Bloeh in Leipzig. 5lit 2 Abbildungen. (Eingegangen am 10. August 1928.) Die Bewegung eines Elektrons im Gitter wird untersucht, indem wir uns dieses.Missing: original paper
  40. [40]
    Semiconductor Band Gaps - HyperPhysics
    Silicon (Si) has a band gap of 1.17 eV at 0K and 1.11 eV at 300K. Germanium (Ge) has 0.74 eV at 0K and 0.66 eV at 300K. InSb has 0.23 eV at 0K and 0.17 eV at ...
  41. [41]
    [PDF] Final Report for Ohio State University Grant DE-FG02-06ER41432 ...
    The most distinctive feature of diamond is its large band gap, 5.5 eV. This large band gap along with the associated large cohesive energy are responsible ...
  42. [42]
    Quantum mechanics of electrons in crystal lattices - Journals
    Szmulowicz F (1999) Kronig - Penney model: a new solution, European Journal of Physics, 10.1088/0143-0807/18/5/015, 18:5, (392-397), Online publication date ...
  43. [43]
    Structure of graphene and its disorders: a review - PMC - NIH
    In monolayer graphene, the conduction band and valence band with zero band gap are formed due to the half-filled π band that permits free-moving electrons.
  44. [44]
    [PDF] Graphite.pdf
    In addition, the delocalization results in loosely bound π-electrons of high mobility, so that the 7-electrons play a dominant role in the electronic properties ...
  45. [45]
    [PDF] Drude, P., 1900, Annalen der Physik 1, 566
    Zur Elektronentheorie der Metalle; von P. Drude. I. Teil. Dass die Elektricitätsleitung der Metalle ihrem Wesen nach nicht allzu verschieden von der der ...
  46. [46]
    Metals: the Drude model of electrical conduction - DoITPoMS
    The Drude model is a simplistic model for conduction. It uses classical mechanics, and treats the solid as a fixed array of ions, with electrons not bound.Missing: source | Show results with:source
  47. [47]
    Semiconductors and Doping – University Physics Volume 3
    Semiconductor p-type doping creates new energy levels just above the valence band. The Hall effect can be used to determine charge, drift velocity, and charge ...
  48. [48]
    Large Anisotropy of the Electrical Conductivity of Graphite - Nature
    The crystal exhibits some remarkable magnetic properties. Whereas its susceptibility along directions in the basal plane is about - 0·5 × 10 -6 per gm., which ...
  49. [49]
    The Polyene Series - CleanEnergyWIKI
    Jun 20, 2011 · The energy of the first optical transition (π to π* or HOMO to LUMO) decreases in conjugated chains approximately as 1/n where n = number of ...
  50. [50]
    [PDF] Exploring the origin of high optical absorption in conjugated polymers
    High optical absorption in conjugated polymers is explained by the high persistence length of the polymer, which is related to its backbone structure.
  51. [51]
    [PDF] Chapter 1 ELECTROMAGNETICS OF METALS - AMOLF
    Experimentally, the plasma frequency of metals typically is determined via electron loss spectroscopy experiments, where electrons are passed through thin ...
  52. [52]
    Plasma Frequency - an overview | ScienceDirect Topics
    Metals are opaque and highly reflective below the plasma frequency, while they become transparent in the ultraviolet.Missing: delocalized | Show results with:delocalized
  53. [53]
    [PDF] The Lorentz Oscillator and its Applications - MIT OpenCourseWare
    The plasma frequency for “non-plasma” materials stands for the natural collective oscillation frequency of the “sea” of electrons in the material, not of ...
  54. [54]
    Electronic Spectroscopy - Interpretation - Chemistry LibreTexts
    Jan 29, 2023 · The Laporte forbidden (symmetry forbidden) d-d transitions are shown as less intense since they are only allowed via vibronic coupling. In ...
  55. [55]
    Role of Vibronic Coupling for the Dynamics of Intersystem Crossing ...
    Mar 7, 2025 · The triplet-singlet radiative transition in aromatic hydrocarbons is the most forbidden electronic transition known for polyat. mols. It is ...
  56. [56]
    [PDF] Physics of Optoelectronic Devices - Light-Emitting Diodes - Vishay
    In order to permit effective recombination, the materials should have direct band-to-band transitions or permit other recombination paths with high efficiency.
  57. [57]
    Effect of Extending the Conjugation of Dye Molecules on the ...
    It is worthy to note that A6 shows a bathochromic shift of the ICT band and obtains the widest spectra response, indicating more interaction between the ...
  58. [58]
    Bathochromic Effect - an overview | ScienceDirect Topics
    The bathochromic effect refers to the phenomenon where the addition of substituents to an aromatic ring, such as in benzene derivatives, results in a shift ...